Noaa 34225 DS1

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Journal of

Plankton Research academic.oup.com/plankt

J. Plankton Res. (2021) 1–15. doi:10.1093/plankt/fbab008

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


BLOOFINZ - Gulf of Mexico
Mesozooplankton biomass, grazing
and trophic structure in the bluefin
tuna spawning area of the oceanic Gulf
of Mexico
MICHAEL R. LANDRY* AND RASMUS SWALETHORP
scripps institution of oceanography, university of california, san diego, 9500 gilman dr., la jolla, ca 92093-0227, usa

*corresponding author: mlandry@ucsd.edu

Received October 13, 2020; editorial decision January 13, 2021; accepted February 1, 2021

Corresponding editor: Xabier Irigoyen

We investigated size-fractioned biomass, isotopes and grazing of mesozooplankton communities in the larval habitat
of Atlantic bluefin tuna (ABT) in the oceanic Gulf of Mexico (GoM) during the peak spawning month of May.
Euphotic-zone biomass ranged from 101 to 513 mg C m−2 during the day and 216 to 798 mg C m−2 at
night. Grazing varied from 0.1 to 1.0 mg Chla m−2 d−1 , averaging 1–3% of phytoplankton Chla consumed d−1 .
Carnivorous taxa dominated the biomass of > 1-mm zooplankton (78% day; 60% night), while only 13% of
smaller zooplankton were carnivores. δ 15 N enrichment between small and large sizes indicates a 0.5–0.6 trophic-
step difference. Although characteristics of GoM zooplankton are generally similar to those of remote oligotrophic
subtropical regions, zooplankton stocks in the ABT larval habitat are disproportionately high relative to primary
production, compared with HOT and BATS averages. Growth-grazing balances for phytoplankton were resolved
with a statistically insignificant residual, and trophic fluxes from local productivity were sufficient to satisfy C demand
of suspension feeding mesozooplankton. While carnivore C demand was met by local processes in the central GoM,
experiments closer to the coastal margin suggest the need for a lateral subsidy of zooplankton biomass to the oceanic
region.

KEYWORDS: zooplankton; biomass; subtropical; oligotrophic; growth-grazing balance; production; carbon


demand; suspension feeders; carnivores; migrants; active export

available online at academic.oup.com/plankt


© The Author(s) 2021. Published by Oxford University Press. All rights reserved. For permissions, please e-mail: journals.permissions@oup.com
JOURNAL OF PLANKTON RESEARCH VOLUME 00 NUMBER 00 PAGES 1–15 2021

INTRODUCTION during the peak spawning month of May. The use of


similar methodology allows us compare zooplankton and
The central role of mesozooplankton as trophic inter-
environmental characteristics in the GoM to those at
mediaries between primary producers and fish is well
oligotrophic subtropical sites in the Atlantic and Pacific
recognized in the classic paradigm of marine food webs
Oceans that have been systematically studied for decades,
(Hardy, 1924) and in studies of larval fish ecology, where
and to process studies in other warm-water open-ocean
the temporal matching of prey resources to the needs
systems. We further draw upon results from companion
of larvae in seasonally dynamic environments is often

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


studies of phytoplankton production (Yingling et al., this
viewed as a critical determinant of recruitment success
issue) and phytoplankton growth and microzooplankton
(Hjort, 1914; Cushing, 1990). In less dynamic larval nurs-
grazing (Landry et al., this issue) to put zooplankton
ery areas, such as oligotrophic subtropical waters, rela-
biomass and grazing in the context of broader trophic
tionships between fish larvae and zooplankton prey are
fluxes. From these, we address the questions: How much
likely just as important, though more subtle and nuanced
does mesozooplankton grazing contribute to the balance
(Llopiz et al., 2014; Landry et al., 2019). For example, like
of phytoplankton growth rate and fates in GoM larval
many tuna and billfish species, Atlantic bluefin tuna (ABT,
ABT habitat? Are trophic fluxes in these waters sufficient
Thunnus thynnus) spawn in nutritionally dilute subtropical
to satisfy the carbon demands of actively growing
waters rather than in more-productive adjacent coastal
zooplankton? What are the relative magnitudes of trophic
margins as an apparent tradeoff between larval starvation
flows to mesozooplankton in these areas?
and predation risks (Bakun and Broad, 2003; Bakun,
2013; Shropshire et al., this issue). Further complicating
interpretation of this strategy, however, ABT larvae are
concentrated in fronts and eddies rather than randomly METHODS
distributed (Bakun, 2006, 2013; Alemany et al., 2010; Lin-
do-Atichati et al., 2012), suggesting that specific properties Zooplankton sample collection
of mesoscale habitats might matter more to larval success Mesozooplankton samples were taken in offshore olig-
than temporal dynamics or mean open-ocean conditions. otrophic waters of the GoM during BLOOFINZ-GoM
Extensive sampling of ABT larvae in the Gulf of Mex- cruises NF1704 (7 May–2 June 2017) and NF1802
ico (GoM) since the 1970s has defined narrow environ- (27 April–20 May 2018) on NOAA Ship Nancy Foster.
mental windows in which ABT can be found in oceanic Sampling was guided by the habitat index model of
waters of the GoM, as well as their particular prevalence Domingues et al. (2016) based on real-time satellite
in the outer rings of anticyclonic loop eddies that spin off imagery (Gerard et al., this issue). Over the two cruises,
periodically from the Loop Current (Muhling et al., 2010, five quasi-Lagrangian process studies, hereafter “cycles”
2011; Lindo-Atichati et al., 2012). Domingues et al. (2016) C1–C5, of 2–4-day duration were conducted following
linked these relationships to satellite measurements of the paths of a satellite-tracked drifter with a mixed-
seasurface height and temperature, establishing a dimen- layer drogue (Fig. 1; Landry et al., 2009). Substantial
sionless index that both successfully captured the general abundances of ABT larvae were found in mixed-layer
features of larval spatial distributions and temporal vari- (upper 25 m) waters during C1 (2017) and C5 (2018)
ability in the GoM and also explained 58% of interannual (Shiroza et al., this issue). During cycle experiments,
stock recruitment variability over a two-decade period we sampled each mid-day (1100–1300) and mid-night
(1993–2011). The trophic ecology of ABT larval habit, (23:00–01:00) with a 1-m diameter ring net (0.2-mm
however, is entirely unexplored. Coherent plankton food- Nitex mesh) towed obliquely through the euphotic zone
web studies have not been done in any oceanic waters of at a ship speed of ∼1.5 kt. Additional transect samples
the GoM, and there are no prior estimates of mesozoo- (T1 and T2) were taken across frontal features at the
plankton carbon biomass or community grazing in the beginning of the 2018 cruise at irregular times of day
region. The present study thus addresses a void in basic (mostly afternoons to evenings between 13:30 and 20:20).
knowledge of regional ecology while also contributing to Tow depth was controlled directly by shipboard readout
a broader system-level process investigation of ABT larval from a pressure sensor at the point of net attachment
habitat in the oceanic GoM (Gerard et al., this issue). on the hydrowire. Tow distance and volume filtered were
As part of the BLOOFINZ-GoM project (Bluefin measured with a calibrated General Oceanics flowmeter
Larvae in Oligotrophic Ocean Foodwebs, Investigation (GO, Miami, FL) attached across the net mouth, assuming
of Nitrogen to Zooplankton in the GoM), we investi- 100% filtration efficiency. For NF1704, pressure rating of
gated biomass and grazing rates of mesozooplankton the hydrowire depth sensor limited net tows to 100-m
communities in ABT larval habitat in 2017 and 2018 depth. For NF1802, mean tow depth was 133 ± 2 m.

2
LANDRY AND SWALETHORP MESOZOOPLANKTON BIOMASS, GRAZING AND TROPHIC STRUCTURE

and depth of tow, and fraction of sample analyzed. The


remaining dried sample was subsequently scraped off the
filter, ground to a powder and subsampled by weight for
carbon (C), nitrogen (N) and stable isotope (13 C and 15 N)
analyses. Individual pyrosomes were treated similarly to
the size-fractioned samples by first measuring total WW
and dividing the WW into subsamples for pigment, DW

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


and CN analyses.
CN subsamples were weighed in small tin boats, packed
into pellets, and analyzed by standard elemental analyzer,
isotope ratio mass spectrometry (EA-IRMS) (Owens and
Rees, 1989) at the Isotope Biogeochemistry lab at Scripps
Institution of Oceanography. The continuous flow system
consisted of a Perkin Elmer CHN analyzer coupled to
a Thermo/Finnigan Delta Plus IRMS. Acetanilide was
Fig. 1. Sampling locations for mesozooplankton day and night net tows
in the GoM. Traces in map inlay depict drift trajectories of Cycles 1–3 the standard used for measurement stability on every run.
in May 2017 and C4–5 in May 2018. Additional transect sampling (T1 C and N biomass estimates (mg m−2 ) were computed for
and T2) was done in 2018. each size fraction from C:DW and N:DW ratios. Stable
isotope values are reported in standard δ () notation
relative to atmospheric N2 and Vienna Pee Dee Belemnite
Upon retrieval, cod-end contents were anesthetized for carbon.
with CO2 (ice-cold soda water; Kleppel and Pieper 1984)
to slow metabolism and gut evacuation and immediately
split with a Folsom plankton splitter. Half of the tow was Gut pigment and grazing estimates
preserved in borate-buffered 5% formalin. The remain- WW subsamples were placed in glass tubes with 7 mL
ing half was size fractioned by wet sieving through nested of 90% acetone and homogenized (multiple 20-s bursts)
screens of 5, 2, 1, 0.5 and 0.2-mm Nitex mesh to produce in an ice bath with a Vibracell sonicator probe. They
5 size classes of 0.2–0.5, 0.5–1, 1–2, 2–5 and 5+ mm. were then extracted overnight (18–24 h) in a −20◦ C
Organisms retained on each mesh were concentrated freezer and warmed to room temperature in a dark con-
onto separate pre-weighed filters (47-mm diameter) of tainer prior to analysis. The homogenate was shaken and
0.2-mm Nitex, rinsed with isotonic ammonium formate centrifuged (5 min at 3000 rpm) to remove particulates.
solution to remove interstitial sea salt, then placed in indi- Concentrations of chlorophyll a (Chla) and phaeopig-
vidual petri dishes and frozen at −80◦ C for later analysis. ments (Phaeo) were then measured by the acidification
On several occasions, we collected small pyrosomes (5– method using a 10-AU fluorometer (Strickland and Par-
8 cm) that were too few to subsample quantitatively. We sons 1972). Water-column estimates of depth-integrated
removed these prior to splitting and size fractioning and Chla for the euphotic zone were made similarly from
froze them separately in 50-mL Falcon tubes for weighing analyses of duplicate 0.25 L samples collected from CTD
and gut pigment analyses. hydrocasts, extracted for 24 h in 90% acetone, and mea-
sured on the same fluorometer.
For each size-fraction analyzed, we computed the
Biomass and stable isotope analyses depth-integrated concentration of gut pigment (Chla,
In the laboratory, frozen size-fractioned zooplankton on Phaeo) in the euphotic zone as:
Nitex filters were thawed, set briefly on blotting paper to
remove excess water and weighed moist for total sample pig∗ D
GPC =
wet weight (WW). Wet samples were subsampled for gut vol∗ f
pigments by removing replicate portions of the biomass
and recording weights before and after each subsampling where GPC is gut pigment content (mg m−2 ), pig is the
(fraction of total WW removed). The remaining wet measured pigment value (mg), f is fraction of sample
biomass on the filters was oven dried at 60◦ C for 24 h analyzed, D is depth of tow (m) and vol is the volume of
before weighing dry for the dry weight to wet weight water filtered (m3 ). To be conservative, our zooplankton
ratio (DW:WW). For each size fraction, zooplankton DW grazing estimates are based only on measured gut Phaeo
(mg m−2 ) was calculated from the measured WW (less values, without correction for inefficiencies in converting
initial filter weight), DW:WW ratio, measured volume Chla into Phaeo.

3
JOURNAL OF PLANKTON RESEARCH VOLUME 00 NUMBER 00 PAGES 1–15 2021

We estimated grazing rates (G, mg pigment m−2 time−1 ) Sheader (1997):


for each size fraction and for the total zooplankton assem-      
blage as G = GPC ∗ K , where K (min−1 ) is the gut evac- log10 G d−1 = −0.2962∗ log10 Ci μg C ind−1 + 0.0246∗ T ◦ C –1.1355.
uation rate constant. For K , we used a gut passage rate
of 2.1 h−1 measured under similar surface water temper- For both respiration and growth rate equations, we
atures in the equatorial Pacific (Zhang et al., 1995). To used mean estimates of individual zooplankton carbon
compute carbon-specific rates of phytoplankton grazing (2.4, 7.4, 41, 140 and 2 782 μg C ind−1 for the 0.2–0.5 to
5+ mm size fractions, respectively) determined from mea-

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


by the zooplankton size classes, we divided G by carbon
biomass (mg C m−2 ). Daily removal rates of phytoplank- sured size-fractioned abundances and carbon biomass in
ton by zooplankton grazing were computed as G ∗ Chlz −1 , 144 net tows from the subtropical Pacific (Landry et al.,
where Chlz is the depth-integrated concentration of Chla 2001). Temperature was the mean euphotic-zone value
in the euphotic zone (mg Chla m−2 ). (T EZ ) in Table I. Respiration rates for each size class were
converted to carbon equivalents using the molar volume
of an ideal gas at standard temperature and pressure
Carnivorous feeders (22.4 L mol−1 ), the respiratory quotient (0.97; Hernán-
As a relatively simple index of trophic structure, we dez-León and Ikeda, 2005), and the molecular weight of
divided the mesozooplankton community into two feed- carbon (12 g C mol−1 ) and multiplying by zooplankton
ing groups: animals considered to be exclusively carniv- abundance (total C biomass/C ind−1 ). Growth rates (d−1 )
orous (only consuming other animals) and omnivorous were converted to production estimates by multiplying by
suspension feeders (mixed feeding on phytoplankton, sus- C biomass of the zooplankton assemblage (mg C m−2 ) and
pended particles, heterotrophic protists and likely small averaging day and night estimates at each station. Cal-
developmental stages of other animals). For this, we vol- culations of ingestion rates (carbon demand) needed to
umetrically subsampled the formalin-preserved split of satisfy respiration and growth requirements also assumed
the zooplankton samples with a Stempel pipette after 70% absorption efficiency for ingested food (Steinberg
thorough mixing and size fractioned the sample through and Landry, 2017). Estimates of mesozooplankton con-
a 1-mm Nitex screen. Except for C2, 2-day and 2-night tribution to carbon export by active diel vertical migra-
tows were subsampled from each cycle. After rinsing with tors were similarly calculated from the biomass differ-
fresh water, zooplankton taxa in the > 1-mm fraction ences between paired day and night tows (size-fractioned
were identified under a dissecting microscope and sorted migrant biomass) and the mean temperature in the 300–
into carnivores and suspension feeders. Each group was 500-m depth range assuming that migrants spend 12 h
placed together on pre-weighed Nitex, dried for 24 h at per day in this stratum (Al Mutairi and Landry, 2001).
60◦ C and measured as DW. Percent carnivore was deter-
mined as the fraction of total subsample DW attributable
to carnivorous taxa. For the < 1-mm fraction, we deter- RESULTS
mined percent carnivore in terms of relative abundances
by dispersing the subsample in a Bogorov sorting tray and Environmental conditions
enumerating the first ∼ 275 animals encountered (range All zooplankton net tows were conducted in warm olig-
255–316) into the two feeding categories. otrophic waters with low mixed-layer Chla and prominent
deep chlorophyll maxima (DCM). Mixed-layer depths
(MLD) were generally in the range of 20–35 m but
Metabolic and growth requirements substantially deeper and more variable for T1 sampling
To estimate feeding requirements for well-nourished (56 ± 17 m) and shallowest (12 ± 0.5 m) for C5 (Table I).
zooplankton, we calculated expected rates of respiration DCM depths varied by > 40 m on each cruise but were
and growth from empirical relations based on zooplank- 22-m shallower, on average, during 2018 compared with
ton size and environmental temperature. For respiration, 2017 (p < 0.001). Conversely, integrated concentrations
we used the equation of Ikeda (1985): of euphotic-zone Chla were higher in 2018 than in 2017
for all composite sampling locations, and almost double
    on average (p < 0.001). While mixed-layer Chla varied
ln Ro μl O2 ind−1 h−1 = 0.8354∗ ln Ci mg C ind−1
  slightly, but significantly for 2017 and 2018 (0.11± 0.01
+0.0601∗ T ◦ C + 0.5254,
versus 0.09 ± 0.01 mg Chla m−3 , respectively; p < 0.004),
where Ci is the average carbon content of an individ- most of the difference between years was in the lower
ual zooplankter in size-fraction i. Potential growth rates euphotic zone, as indicated by larger DCM peak concen-
(G, d−1 ) were computed from the equation of Hirst and trations in 2018 (0.71 ± 0.04 versus 0.54 ± 0.04 mg Chla

4
LANDRY AND SWALETHORP MESOZOOPLANKTON BIOMASS, GRAZING AND TROPHIC STRUCTURE

Table I: Environmental conditions for zooplankton net tows in the GoM during May 2017 (Cycles 1–3)
and April–May 2018 (Transects 1–2; Cycles 4–5)
Samples Dates MLD (m) DCM (m) T ML (◦ C) T EZ (◦ C) Chla (mg m−2 )

Cycle 1 11–14 May 31.5 ± 4.0 100 ± 1.3 24.5 ± 0.12 23.6 ± 0.36 17.7 ± 1.0
Cycle 2 16–18 May 24.2 ± 1.1 120 ± 7.9 25.3 ± 0.06 24.4 ± 0.04 16.2 ± 1.4
Cycle 3 27–30 May 25.8 ± 4.1 137 ± 1.7 26.7 ± 0.07 24.8 ± 0.02 13.7 ± 0.7
Trans 1 30 Apr–1 May 56.1 ± 17.2 100 ± 10.8 26.1 ± 0.28 25.2 ± 0.77 35.1 ± 6.7
26.6 ± 0.4 92 ± 4.4 24.9 ± 0.10 24.2 ± 0.49 43.0 ± 0.3

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


Trans 2 2 May
Cycle 4 5–9 May 23.7 ± 1.2 111± 7.1 25.6 ± 0.07 24.8 ± 0.05 24.1 ± 1.2
Cycle 5 15–19 May 12.4 ± 0.5 77 ± 3.9 25.5 ± 0.14 24.2 ± 0.22 29.2 ± 5.4
2017 28.1 ± 1.6 118 ± 4.0 25.5 ± 0.23 24.2 ± 0.15 15.8 ± 0.5
2018 26.4 ± 5.0 95 ± 4.6 25.6 ± 0.10 24.5 ± 0.20 29.4 ± 2.3

MLD (m) = mixed layer depth, where seawater density is 0.1 kg m−3 greater than at 10 m. DCM (m) = depth of the Deep Chlorophyll Maximum;
T ML (◦ C) = mean temperature of the mixed layer; T EZ (◦ C) = mean euphotic-zone temperature to the DCM; and CHLZ (mg m−2 ) is mean
depth-integrated Chla to the DCM. 2017 and 2018 cruise are averages for all tows in cruise years. Uncertainties are standard errors of mean
values.
Bold font is used to highlight the averages of cruise years, as opposed to the averages of individual cycle experiments in the two years.

Table II: Carbon (C), nitrogen (N), WW and DW relationships for mesozooplankton size classes in the
oceanic GoM
Size fraction DW:WW (%) C:DW (%) N:DW (%) C:N

0.2–0.5 mm 14.8 ± 0.8 36.8 ± 0.3 8.18 ± 0.09 4.51 ± 0.03


0.5–1 mm 16.5 ± 1.0 39.4 ± 0.3 9.91 ± 0.08 4.28 ± 0.02
1–2 mm 11.6 ± 0.4 38.5 ± 0.3 9.06 ± 0.08 4.25 ± 0.03
2–5 mm 13.4 ± 1.8 35.7 ± 0.4 8.61 ± 0.11 4.15 ± 0.04
5+ mm 8.5 ± 2.0 26.5 ± 1.6 6.21 ± 0.45 4.21 ± 0.16
Total 12.1 ± 0.7 36.0 ± 1.6 8.47 ± 0.14 4.25 ± 0.03

Averages are based on 44 day and nighttime euphotic-zone net tows in May 2017 and May 2018. Uncertainties are standard errors of mean
values.
Bold font is used to highlight the averages of cruise years, as opposed to the averages of individual cycle experiments in the two years.

m−3 ; p < 0.001). Despite the Chla differences, water tem- (36.8%; P < 0.001) > 2–5 mm (35.7%; P < 0.05) > 5+
peratures varied narrowly ∼24.5–26.5◦ C for the mixed mm (26.5%; P < 0.001; Table II). The C:N ratio for the
layer and 23–25◦ C for the euphotic zone and were not 0.2–0.5 mm fraction (4.51) is significantly higher than
different, on average, between years (p > 0.20). both of the 1–2 and 2–5 mm fractions (4.25 and 4.15,
respectively; P < 0.0001), while all other size fractions
show no C:N differences. The difference in the small-
Elemental and weight relationships est size fraction may reflect some contamination from
Day-night differences in DW:WW ratios were not large phytoplankton or detritus, which have higher C:N
significant for any of the zooplankton size classes compositions than zooplankton.
(P > 0.25). In 2017, C:DW was significantly lower at night
(33.4 ± 1.9%) than during the day (38.4 ± 0.6%) due to
more large gelatinous animals in the euphotic zone at Biomass variability
night, which lowered the carbon content of the 5+ mm Mesozooplankton carbon biomass varies substantially in
size fraction from 30.6 to 19.1% of DW (P < 0.035). our analyses among sampling locations (cycles), between
For 2018 and the other 2017 size fractions, day-night day-night tows, between cruise years and among size
variability in %C:DW was insignificant. Because most classes (Fig. 2). Among cycles, mean day biomasses
of the variability in elemental and weight conversions range 5 fold from 101 ± 12 to 513 ± 71 mg C m−2
occurs among size classes rather than time of day, we for C2 and C5, respectively. Nighttime biomasses
highlight these differences by averaging all day and night range 3.7 fold from 216 ± 16 (C1) to 798 ± 28 (C5)
tows together in Table II. mg C m−2 . Within years, night tow biomass averages
Carbon is the biomass basis for all subsequent anal- are significantly higher than day averages in 2017
yses. The C:DW relationship is highest for the 0.5–1 (239 ± 15 versus 144 ± 16 mg C m−2 ; P < 0.001), but
and 1–2 mm fractions (39.4 and 38.5%, respectively) not in 2018 (514 ± 69 versus 424 ± 53 mg C m−2 ;
and decreases significantly in the order of 0.2–0.5 mm P > 0.31). Between years, both day and night biomass

5
JOURNAL OF PLANKTON RESEARCH VOLUME 00 NUMBER 00 PAGES 1–15 2021

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


Fig. 2. Size-fractioned carbon biomass structure of mesozooplankton Fig. 3. Size-fractioned grazing of mesozooplankton in the GoM during
in the GoM during Cycles 1–5 drift experiments and total average Cycles 1–5 drift experiments and average total grazing for day (D) and
carbon for day (D) and night (N) tows in 2017 and 2018 cruises. Totals night (N) tows in 2017 and 2018 cruises. Uncertainties are standard
for 2018 include T1 and T2 sampling, in addition to C4 and C5. errors of mean estimates.
Uncertainties are standard errors of mean estimates.

averages are significantly higher in 2018 than 2017 2–5 mm fraction has a marginal difference between night
for the total (P < 0.002) and for each individual size and day tows (P < 0.048). These departures from the
fraction in day and night tows (P < 0.02), except for biomass trends reflect differences among size classes and
5+ mm animals during the day (P > 0.40). Among cycles in carbon-specific grazing rates (Table III).
size classes for all net tows, the 1–2 mm fraction Overall, we found no significant differences between
contributes most to biomass (29.9 ± 1.3%; P < 0.016), day and night estimates of C-specific grazing for any of
the 0.5–1 and 2–5 mm fractions contribute comparably the size fractions (P > 0.05 for 5+ mm; P > 0.21 for all
(23.7 ± 1.2 and 24.8 ± 1.6%, respectively; P > 0.60), other sizes) or for the whole community (P > 0.38). C-
and further contributions decline in order of 0.2– specific grazing does, however, vary significantly among
0.5 mm (16.5 ± 0.9%; P < 0.001) and 5+ mm fractions all size classes, as might be expected, in exact inverse
(5.0 ± 0.8%; P < 0.001). For night tows only, inclusive order of mean animal size (Table III, P < 0.03:0.2–
of diel migrants, the 2–5 and 1–2 mm fractions are 0.5 mm (2.43 ± 0.23μg Chl mg C d−1 ) > 0.5–1 mm
the biomass co-dominants (29.1 ± 2.1 and 27.4 ± 1.7%, (1.65 ± 0.12 μg Chl mg C d−1 ; P < 0.003) > 1–2 mm
respectively; P > 0.54), and further contributions decline (1.14 ± 0.07 μg Chl mg C d−1 ; P < 0.001) > 2–5 mm
in order of 0.5–1 mm (22.2 ± 1.6%; P < 0.03), 0.2– (0.85 ± 0.02 μg Chl mg C d−1 ; P < 0.03) > 5+ mm
0.5 mm (14.2 ± 1.1%; P < 0.001) and 5+ mm fractions (0.05 ± 0.02 μg Chl mg C d−1 ; P < 0.001). Although C-
(7.2 ± 1.2%; P < 0.001). specific grazing estimates are higher in 2017 than 2018
across all sizes (Table III), the differences are significant
for only the 0.2–0.5 mm (P < 0.004), 2–5 mm (P < 0.016)
Grazing rates and size relationships and total (P < 0.001) size categories.
Mean grazing rate estimates range from 0.1 to 1.0 mg
Chla m−2 d−1 for the whole mesozooplankton community
and show substantial variability among cycles, day and Isotopic composition
night tows, years and size classes (Fig. 3). The main dif- Day and night tows did not give statistically different
ferences are shaped by the variability in biomass (Fig. 2). isotopic values for any of the zooplankton size fractions
Similar to biomass trends, for example, the lowest and and are reported as mean values in Fig. 5. δ 15 N also did
highest grazing rates are for C2 and C5 (0.22 ± 0.08 and not differ among size fractions in the 2 years, but δ 13 C
0.82 ± 0.13 mg Chla m−2 d−1 , respectively; P < 0.004), was 0.3–0.6 lower (P < 0.01) for the smallest 3 frac-
and 2018 grazing is higher overall compared with tions (0.2–2 mm) in 2018 compared with 2017. Among
2017 (0.52 ± 0.08 and 0.33 ± 0.03 mg Chla m−2 d−1 ; size classes, δ 15 N increases significantly from 0.2–0.5 mm
P < 0.025). In contrast to biomass, however, nighttime (3.38 ± 0.19) to 0.5–1 mm (4.05 ± 0.19; P < 0.014)
grazing estimates do not differ significantly from day and to 1–2 mm (4.96 ± 0.22; P < 0.003) zooplankton,
estimates for the full community for either of the 2 years but size fractions > 1 mm have similar values in the range
separately or both years combined (0.48 ± 0.07 versus of 5.0–5.25 (P > 0.35) (Fig. 5A). Assuming that a 3.4
0.39 ± 0.06 mg Chla m−2 d−1 , P > 0.38), and only the increase in bulk δ 15 N corresponds to one full trophic level

6
LANDRY AND SWALETHORP MESOZOOPLANKTON BIOMASS, GRAZING AND TROPHIC STRUCTURE

Table III: Carbon-specific grazing rates for size-fractioned mesozooplankton in the GoM during Cycle 1–5
experiments and for all tows in 2017 and 2018 cruises
Carbon-specific grazing rates (μg Chla mg C−1 d−1 )

Samples 0.2–0.5 mm 0.5–1 mm 1–2 mm 2–5 mm 5+ mm Total

Cycle 1 2.51 ± 0.22 1.68 ± 0.18 1.06 ± 0.15 0.85 ± 0.27 — 1.50 ± 0.13
Cycle 2 2.19 ± 0.07 1.45 ± 0.14 0.81 ± 0.18 1.05 ± 0.45 0.14 ± 0.14 1.22 ± 0.12

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


Cycle 3 4.25 ± 0.64 2.28 ± 0.17 1.53 ± 0.13 1.50 ± 0.33 0.12 ± 0.08 2.06 ± 0.17
Cycle 4 1.67 ± 0.32 1.09 ± 0.11 0.90 ± 0.09 0.39 ± 0.11 0.06 ± 0.06 0.81 ± 0.11
Cycle 5 2.04 ± 0.31 1.84 ± 0.39 1.35 ± 0.22 0.88 ± 0.18 0.001 ± 0.001 1.24 ± 0.15
2017 3.18 ± 0.35 1.88 ± 0.13 1.20 ± 0.10 1.17 ± 0.20 0.08 ± 0.04 1.67 ± 0.12
2018 1.86 ± 0.26 1.47 ± 0.17 1.09 ± 0.10 0.61 ± 0.09 0.03 ± 0.20 1.05 ± 0.09

Uncertainties are standard errors of mean values.


Bold font is used to highlight the averages of cruise years, as opposed to the averages of individual cycle experiments in the two years.

of isotopic enrichment (Minawaga and Wada, 1984), the Carbon demand and migrant export
1.87 difference between smaller and larger zooplank- Because temperature varies little among the sampling
ton size fractions suggests a 0.55 step increase in mean locations (Table I), calculated estimates of mesozooplank-
trophic position. ton carbon demand (Fig. 6) closely follow biomass vari-
Despite the slight negative trend, none of the size- ability (Fig. 2), except that smaller size classes contribute
fraction comparisons for δ 13 C are significant (P > 0.35; proportionally more to total C demand because of the
Fig. 5B). On average, therefore, zooplankton in the differ- size functions in respiration and growth equations (Ikeda,
ent size classes appears to share a common food-web base. 1985; Hirst and Sheader, 1997). C demand ranges from
67.5 ± 0.5 to 15.1 ± 0.1% body C d−1 for 0.2–0.5 mm
Carnivorous feeders and 5+ mm size fractions, respectively, and averages
Among smaller (0.2–1 mm) mesozooplankton, carnivo- 42 ± 0.7% body C d−1 for the community. Day-night
rous taxa comprised a small, but relatively consistent, and cruise differences of C demand as % body C d−1
fraction of the community (13.1 ± 0.7%) that did not vary are insignificant. The total feeding required to meet C
significantly (P > 0.07) between day and night sample demand is, however, significantly higher in 2018 than
collections (Fig. 5). Copepods of the family Corycaeidae 2017 (194 ± 17 versus 80 ± 6 mg C m−2 d−1 ; P < 0.001).
and small chaetognaths were the main contributors to this The annual averages in Fig. 6 separate portions of
feeding group in the samples examined. total C demand that can be satisfied by carnivorous
Among larger (>1 mm) zooplankton, carnivorous taxa versus omnivorous suspension feeding based on the con-
dominated euphotic-zone biomass, significantly more tributions of those feeding groups to small (0.2–1 mm)
during day (78.4 ± 5.0%) than night tows (60.5 ± 4.5%; and large (>1 mm) zooplankton in Fig. 5. The propor-
P < 0.013) (Fig. 5). This difference reflects the entry tions of C demand assigned to each feeding group were
of diel vertical migrants into the euphotic zone at partitioned according to average % carnivore for day
night, especially large euphausiids, which we scored and night tows of each cycle (T1 and T2 samples were
conservatively as omnivorous suspension feeders given portioned according to mean day and night estimates
the undocumented feeding status of most species. for C4 and C5 combined), then summed for the total
Conspicuous taxa comprising the carnivorous group were community and averaged per year. Of the total C demand
abundant large chaetognaths and large copepods of the for 2017, 29.1 ± 2.1 and 50.4 ± 4.0 mg C m−2 d−1 are
genus Euchaeta, most of which appeared to reside in the ascribed to carnivores and suspension feeders, respec-
euphotic-zone day and night. In addition to euphausiids, tively. For 2018, C demand divides into 74.9 ± 7.0 and
copepods of the genera Neocalanus and Eucalanus were 119.2 ± 10.8 mg C m−2 d−1 for carnivores and suspension
significant contributors to suspension feeding biomass in feeders. On average, ∼ 38% of total C demand can be
most samples, while salps were relatively small and rare. satisfied by direct predation on metazooplankton.
This analysis does not account for the pyrosomes removed Mean temperature in the 300–500 m depth stratum
from samples prior to splitting, which were few in number ranged from 10.6 to 12.4◦ C (11.7 ± 0.02◦ C) for the 5
(5 small colonies in 44 tows) and only collected at night. cycles where we had paired day and night tows to com-
Considering pyrosomes would therefore have lowered the pute active export flux by respiration of diel migrants.
nighttime percentage of carnivorous biomass, though not Some of the estimates were slightly negative when day
substantially. biomass values exceeded nighttime values (range − 3.3 to

7
JOURNAL OF PLANKTON RESEARCH VOLUME 00 NUMBER 00 PAGES 1–15 2021

15.6 mg C m−2 d−1 ). The mean active export estimate of have been useful in defining broad-scale patterns of
3.7 ± 1.1 mg C m−2 d−1 represents, on average, the daily zooplankton biomass (Moriarty and O’Brien, 2013;
respiratory loss of 3.2 ± 0.4% of the body C of migrant Moriarty et al., 2013). In general, subtropical waters (15–
zooplankton. 40◦ ) are biomass minima between higher zooplankton
stocks in tropical (15◦ S–15◦ N) and temperate-to-high-
latitude (40–90◦ ) waters. We examine the low-latitude,
DISCUSSION
warm-water component of this pattern in Table IV for a

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


Although we intended to sample consistently in this study subset of studies with methods comparable to the present
to depths that included the DCM, that was not possible for study, including time-series sampling in the subtropical
all tows on the 2017 cruise. Because DCMs were deepest North Pacific (Stn. ALOHA, Hawaii Ocean Time-series,
for C2 and C3, the samples collected for those cycles HOT) and subtropical North Atlantic (Sargasso Sea,
were the most compromised by this issue, with missed Bermuda Atlantic Time-series Study, BATS) as well as
DCM sampling potentially affecting estimates of total process studies in the eastern tropical Pacific, equatorial
biomass m−2 , day-night differences, relative size structure Pacific and eastern Indian Ocean.
and grazing rates relative to water-column Chla. For As in the global trend, mesozooplankton biomass is
example, the daytime biomass estimates for C2 and C3 higher in our subset of studies for tropical regions than
are the lowest for all tows while nighttime estimates are adjacent subtropical systems (Table IV). For example,
comparable or higher than for C1 (Fig. 2). Similarly, day- biomass in the central equatorial Pacific (994 ± 77 mg C m−2 ;
time grazing estimates are extremely low for C2, though Décima et al., 2011) is almost three times higher than
not for C3, whereas nighttime grazing for C2 and C3 are the long-term average at Stn. ALOHA in the North
both comparable and higher than C1 and C4 (Fig. 3). Pacific subtropical gyre (349 ± 26 mg C m−2 ; Valencia
A substantial portion of the zooplankton community et al., 2018). Biomass in tropical waters of the eastern
appears to reside in the lower euphotic zone, between Indian Ocean along 110◦ E is also more than double
100 m and the DCM, during the day, but comes into the average in subtropical Indian Ocean Central Water
the upper 100 m at night, along with regular migrants directly to the south (455 ± 18 versus 209 ± 32 mg C m−2 ;
that completely leave the euphotic zone during the day. Landry et al., 2020a). Similarly, various crossing of
Thus, while C2 and C3 euphotic depths were under- the Atlantic Meridional Transect have documented
sampled also at night, nighttime biomass and grazing substantially higher (1.5-16X) zooplankton biomass in
estimates to 100 m provide more useful comparisons the equatorial Atlantic than in subtropical gyres of the
to other cycles. Based on them, the biomass difference North and South Atlantic (Isla et al., 2004; López and
between years appears to be valid, scaling approximately Anadón, 2008; Calbet et al., 2009). Such differences
with the 2-fold higher concentrations of integrated Chla arise because these tropical systems are associated with
in 2018 compared with 2017 (Table I). Grazing, however, divergence or mixing centers that enhance productivity
does not scale proportionately with the Chla difference relative to the more pervasive downwelling of subtropical
between years because the C-specific grazing rate for gyres. Nonetheless, as shown in Table IV, this occurs
the zooplankton community was ∼ 60% higher in 2017 without much of a difference in total integrated Chla
(Table III). because the shallow concentrated euphotic zones of
Due to the sampling depth irregularities for C2 and C3, richer systems are compensated for by deeper euphotic
we focus most of the discussion below on C1, C4 and C5, zones with pronounced DCMs in the subtropics. The
which are also the better characterized experiments for importance of relative productivity, rather than latitude,
phytoplankton growth rates and microzooplankton graz- in driving such differences is illustrated by sampling along
ing (Landry et al., this issue), allowing more comprehen- 160◦ E in the western Pacific (Sun and Wang, 2017).
sive assessments of trophic interactions for these cycles. In In this case, equatorial waters are a local minimum in
the sections below, we first compare results for the GoM mesozooplankton biomass compared with the latitudinal
to other warm-water oceanic regions investigated with trend to the north because they reside in a deep layer
similar methods. We then address the two major questions of oligotrophic waters, the Western Warm Pool, which
of our study. What is the role of mesozooplankton grazing does not allow significant upwelling of nutrients despite
in the balance of phytoplankton growth and losses in the conducive winds.
GoM? How do various trophic processes contribute to GoM results are consistent with the lower primary
satisfying mesozooplankton carbon demand? production and zooplankton biomass expected for
subtropical regions (Table IV). Our mesozooplankton
Comparisons to other open-ocean regions biomass estimates are almost identical to the long-term
Despite numerous corrections for net type, mesh size, average from the subtropical Pacific at Stn. ALOHA (349
tow depth and measurement units, global analyses versus 351 mg C m−2 , respectively), though substantially

8
LANDRY AND SWALETHORP MESOZOOPLANKTON BIOMASS, GRAZING AND TROPHIC STRUCTURE

Table IV: Comparison of environmental conditions, mesozooplankton biomass and grazing estimates
among open-ocean tropical and subtropical regions
T EZ Intgr Chla PrimProd Biomass Grazing
Region Latitude (◦ C) (mg m−2 ) (mg C m−2 d−1 ) (mg C m−2 ) (% Chla d−1 ) Refs

Costa Rica Dome 9–11 ◦N 22.0 ± 0.3 24.1 ± 1.5 1 025 ± 113 1870 ± 72 54 ± 11 1
Equatorial Pacific 4◦ S-4◦ N 24.2 ± 0.2 26.7 ± 0.7 672 ± 3.7 994 ± 77 14 ± 6 2
Tropical E. Indian 11–14◦ S 26.9 ± 0.6 25.7 ± 2.9 ND 455 ± 18 21 ± 4 3
27–35◦ S 19.6 ± 0.4 27.2 ± 3.8 209 ± 32 3 ± 0.4

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


Subtrop. E. Indian ND 3
W. Equat. Pac, 160◦ E 5◦ S–10◦ N 30.0 25.5 ND 259 ND 4
Subtrop. N. Pacific 22–23◦ N 24.4 ± 0.1 23.0 ± 0.2 516 ± 11 349 ± 26 ND 5
Subtrop. Atlantic 31–32◦ N 21.3 ± 0.1 24.8 ± 2.4 455 ± 11 218 ± 4 3 to 24 6
Gulf of Mexico 25.4–27◦ N 24.5 ± 0.2 22.8 ± 0.2 325 ± 14 351 ± 33 2 ± 0.2 7

References: (1) Décima et al. (2016), Landry et al. (2016), Taylor et al. (2016); (2) Balch et al. (2011), Décima et al. (2011), Landry et al. (2011),
Taylor et al. (2011); (3) Landry et al. (2020b); (4) Sun and Wang (2017); (5) Hawaii Ocean Time series (HOT), https://hahana.soest.hawaii.edu/hot/
hot-dogs/, Valencia et al. (2018); (6) Bermuda Atlantic Time Series (BATS), http://bats.bios.edu/data/, Madin et al. (2001), Roman et al. (1993); (7)
This study, Selph et al. (in press), Yingling et al. (this issue).
T EZ (◦ C) and Intgr Chla (mg m−2 ) are mean temperature and depth-integrated Chla of the euphotic zone. PrimProd is the measured rate of
daily primary production from 14 C or 13 C-labeled bicarbonate uptake (mg C m−2 d−1 ). Biomass (mg C m−2 ) estimates are for mesozooplankton
collected in oblique tows with 200-μm mesh nets covering the euphotic zone, originally presented as carbon or calculated from DW using a
C:DW ratio of 0.36. Grazing impact estimates are for the percent of total euphotic-zone Chla (phytoplankton) cleared by mesozooplankton per
day. Uncertainties are standard errors of mean values.

higher than the average for BATS in the Sargasso study, are higher in tropical regions (14–54% Chla d−1 )
Sea (218 ± 4 mg C m−2 ). Though sampled similarly, compared with estimates (2–3% Chla d−1 ) from subtropi-
HOT zooplankton biomass has long been known to cal areas. In this regard, the lower estimate of 3% d−1 for
exceed that at BATS (Roman et al., 2002), and they the subtropical Atlantic is from a study by Roman et al.
remain consistently separated even as biomass has (1993) that compared zooplankton grazing during stable
increased by 60–80% in both systems over the past stratified conditions (August), which prevail for most of
three decades (Sheridan and Landry, 2004; Steinberg the year, to the peak of a late-winter overturn event
et al., 2012; Valencia et al., 2016). Although mean (March–April), which gave the higher (24% d−1 ) grazing
primary production is also higher at Stn. ALOHA (516 impact. The phytoplankton bloom that develops from the
versus 455 mg C m−2 d−1 ), the two sites have different upward mixing of deep nutrients is a regular feature of
biomass:production ratios of 0.68 (HOT) and 0.48 mg the seasonal dynamics of the Sargasso Sea that makes
C (mg C)−1 d−1 (BATS). Paradoxically, relatively high this region more like a tropical upwelling system for a
zooplankton biomass in the GoM is associated with low small portion of the year, as the higher grazing estimate
primary production (325 mg C m−2 d−1 ; Yingling et al., indicates. The extent to which similar seasonal events
this issue), giving a much higher biomass:production ratio might also occur in the GoM is not known, although the
(1.07 mg C (mg C)−1 d−1 ) than either HOT or BATS. One regular passage of strong tropical storms and hurricanes
explanation for this difference is temporal aliasing, as the through the region in late summer have the potential to
GoM samples were only collected in May of 2 years, while generate strong mixing on local scales (Babin et al., 2004;
the time-series station averages are for all months over Avila-Alonso et al., 2020). At least for May, the region was
many years. May sampling may simply catch a seasonal strongly stratified and consistent with other subtropical
imbalance in the magnitude of accumulated zooplankton systems during stratified conditions in showing a relatively
relative to contemporaneous primary production. In low grazing impact of mesozooplankton.
addition, since our GoM sampling was site selective, Several previous studies have commented upon the
rather than spatially random, the difference could reflect relative uniformity of biomass distribution across size
unique aspects of the mesoscale features where ABT classes < 5 mm in subtropical regions of the Atlantic,
larvae are more abundant. Either way, during the ABT Pacific and Indian Oceans (Landry et al., 2001; Madin
spawning peak in the GoM, the ratio of zooplankton et al., 2001; Valencia et al., 2018; Landry et al., 2020a).
biomass to primary production in the larval rearing area In our GoM samples, however, biomass structure was
is substantially higher than would be predicted from the tilted to larger sizes, with the 2–5 mm fraction averaging
averages of well-studied oligotrophic subtropical regions 27% of biomass compared with 16% for 0.2–0.5-mm
of the oceans. animals, whereas these two fractions are similar (20–
Along with the biomass differences, zooplankton graz- 23%) in other regions. Larger size classes contribute
ing impacts on phytoplankton, measured similarly to this more to diel vertical migration. Our estimates of carbon

9
JOURNAL OF PLANKTON RESEARCH VOLUME 00 NUMBER 00 PAGES 1–15 2021

respiration at depth by diel migrating zooplankton, equatorial Pacific (Landry et al., 2011) and Costa
3.7 ± 1.1 mg C m−2 d−1 , are close to comparably Rica Dome (Landry et al., 2016). In the California
calculated active flux rates for the subtropical Pacific Current, measured rates of phytoplankton growth,
(3.6 ± 0.4 mg C m−2 d−1 , Al-Mutairi and Landry, 2001; microzooplankton grazing and mesozooplankton grazing
updated to 4.2 ± 0.2 mg C m−2 d−1 , Valencia et al., 2018) explained 93% of the variability in rates of phytoplankton
but substantially higher than estimates for the subtropical biomass accumulation along a trophic gradient from
Atlantic (1.2 mg C m−2 d−1 , Steinberg et al., 2000). coastal upwelling to oligotrophic open ocean. In the

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


Steinberg et al. (2012) updated the earlier migrant flux equatorial Pacific and Costa Rica Domes, the dynamics
estimates for BATS to account for the increase in time- were reflected in steady-state balances of growth and
series biomass but did not separate the carbon respiratory grazing with low to negligible unresolved residuals. What
component from the total active export (4.1 ± 0.3) that these previously studied systems have in common are
included organic excretion and fecal pellet transport. phytoplankton concentration maxima in or close to the
Based on the multipliers applied, the updated estimate mixed layer, therefore in the depth stratum tracked by
for BATS is 60% of that at Stn. ALOHA (Valencia drifters with mixed-layer drogues. For the current exper-
et al., 2018), so about 2.5 mg C m−2 d−1 for daytime iments, Landry et al. (this issue) observed near steady-
C respiration at depth. Therefore, consistent with the state mixed-layer Chla, with microzooplankton grazing
differences in total community biomass, the magnitude typically equal or exceeding phytoplankton growth, while
of active migrant flux in the oceanic GoM during May growth and grazing were strongly uncoupled in the mid
is more like the average for the subtropical North Pacific to lower photic zone, accounting for the net changes in
than the western subtropical North Atlantic. ambient Chla. Thus, a major challenge in interpreting
growth-grazing balances in the GoM is the large influence
of the portion of the photic zone that shows significant
Phytoplankton growth and grazing balances net biomass rate of change but is not specifically marked
In addition to mesozooplankton grazing rates, measured or tracked by the drifter. It is thus difficult to know
daily rate profiles of phytoplankton growth, microzoo- whether imbalances for the cycles in Table V are due
plankton grazing and net changes in ambient water- to process measurement issues versus lateral inputs that
column Chla from dilution experiments during the cycle are not independently measured. Advective influences on
experiments (Landry et al., this issue) enable a first-order water-column dynamics might also be exacerbated in the
analysis of phytoplankton growth rates and fates for the oceanic GoM by strong gradients in the surrounding
euphotic zone based on fluorometrically measured Chla coastal margins (Kelly et al., in review) and by our
(Table V). C2 is excluded from this analysis because it experimental site selection in mesoscale features. Given
was a 2-day experiment with only one resolved rate pro- these complications, it may be fortuitous that the process
file. Depth-integrated estimates of phytoplankton growth balances are as well resolved as they appear.
rates range from 0.25 to 0.59 d−1 , and microzooplankton
grazing (0.10–0.42 d−1 ) is the major loss term compared
with mesozooplankton grazing impact (0.01–0.03 d−1 ). Trophic flows, trophic structure
Net change in photic zone Chla varies from a net accumu- and carbon demand
lation rate of +0.16 d−1 for C5 to Chla loss of −0.06 d−1 Direct grazing on phytoplankton is typically too low to
for C3. None of the water columns investigated was support full respiration and growth of mesozooplank-
at steady state. For cycles where we observed net Chla ton in open-ocean ecosystems (Dam et al., 1995; Calbet
accumulation in the photic zone (C1, C4 and C5; Landry et al., 2009). Nonetheless, as recently demonstrated for
et al., this issue), the budgets can be reasonably closed, the equatorial Pacific (Landry et al., 2020b), the broader
between −0.04 and + 0.09 d−1 . Due to the net decline in network of trophic flows—including omnivorous feeding
Chla during C3, a large portion of that balance (0.16 d−1 ) on microzooplankton, carnivory and detritivory—might
is unresolved. Thus, for individual cycles, there is mixed still be able to meet the carbon demands of active zoo-
success in reconciling the balance of growth and loss plankton computed from empirical relationships (Fig. 6).
terms. For the daily resolved rates of all 4 cycles, the In Table VI, we evaluate this potential for the GoM
balance residual (0.05 ± 0.04 d−1 ) is equivalent to ∼ 10% for the three experimental cycles with adequate data for
of mean phytoplankton growth rate but is not significantly carbon fluxes. Carbon-based estimates of phytoplankton
greater than zero (t-test, P > 0.11). grazing (7–20 mg C m−2 d−1 ), from daily mean day-night
Similar analyses have demonstrated tighter bal- values of % Chla d−1 cleared by mesozooplankton and
ances of phytoplankton growth rates and fates in the phytoplankton C values for the photic zone (Selph et al.,
California Current Ecosystem (Landry et al., 2009), in press), account for ∼ 23% of total mesozooplankton C

10
LANDRY AND SWALETHORP MESOZOOPLANKTON BIOMASS, GRAZING AND TROPHIC STRUCTURE

Table V: Phytoplankton growth rates and fates during cycle process experiments in the GoM
Rate (d−1 ) Cycle 1 Cycle 3 Cycle 4 Cycle 5 ALL

Phyto growth 0.44 ± 0.06 0.37 ± 0.07 0.59 ± 0.05 0.50 ± 0.06 0.49 ± 0.03
Microzoo graz 0.34 ± 0.09 0.24 ± 0.06 0.42 ± 0.09 0.35 ± 0.07 0.34 ± 0.04
Mesozoo graz 0.02 ± 0.004 0.03 ± 0.002 0.01 ± 0.001 0.03 ± 0.002 0.02 ± 0.002
Net change 0.09 ± 0.05 −0.06 ± 0.08 0.06 ± 0.03 0.16 ± 0.09 0.07 ± 0.04
Residual −0.0004 ± 0.01 0.17 ± 0.05 0.09 ± 0.08 −0.04 ± 0.10 0.05 ± 0.04

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


Growth-grazing balance residuals are the differences between measured rate estimates for phytoplankton growth and the sum of losses to
micro- and mesozooplankton grazing plus the net rate of biomass change in the euphotic zone during the cycle. All rates are measured by
fluorometric Chla and integrated for the photic zone for daily experiments (Landry et al., this issue). Uncertainties are standard errors of mean
values.
Bold font is used to highlight the averages of cruise years, as opposed to the averages of individual cycle experiments in the two years.

Table VI: Comparisons of nutritional resources relative to mesozooplankton carbon demand for experimen-
tal cycles in the GoM
Variable Cycle 1 Cycle 4 Cycle 5

Nutritional resource (mg C m−2 d−1 )


Phytoplankton grazing 20 ± 6 7±1 20 ± 3
Microzoo production 126 ± 35 116 ± 10 121 ± 10
Mesozoo production 26 ± 3 42 ± 5 77 ± 9
Carbon demand (mg C m−2 d−1 )
Suspension feeders 61 ± 8 91 ± 11 159 ± 24
Carnivorous feeders 27 ± 3 56 ± 8 106 ± 9

Phytoplankton grazing is calculated from % Chla grazed d−1 and phytoplankton C biomass m−2 . Microzoo production is calculated from
the combined C grazing rates of microzooplankton on phytoplankton and bacteria from Landry et al. (this issue), assuming a 30% gross
growth efficiency. Mesozoo Production is computed as described from empirical equations of Hirst and Sheader (1997). Carbon Demands are
computed for suspension feeders and carnivores as described for Fig. 6. All rates are mg C m−2 d−1 . Uncertainties are standard errors of
mean values.

demand for C1 and much lower (5–8%) percentages for deficits of 25–27%. Carnivore C demand could be
C4 and C5. In comparison, microzooplankton produc- overestimated because it derives from a copepod-
tion (116–126 mg C m−2 d−1 ), from daily integrated C- based growth relationship (Hirst and Sheader, 1997),
based grazing rates of microzooplankton on phytoplank- whereas the GoM carnivore biomass is dominated by
ton and bacteria (Landry et al., this issue) and assuming chaetognaths. Indeed, the multiple regression equation
a gross growth efficiency of 30% (Straile, 1997), is 6–17 developed by Ikeda (2014) for respiration of diverse
times more important as a potential food resource than zooplankton communities has a negative coefficient,
phytoplankton. For C1 and C4, the combined resources relative to copepods, for the contribution of chaetognaths
from phytoplankton and microzooplankton substantially compared with a positive coefficient for euphausiids. For
exceed the estimated C demands of suspension feeders. our samples, overestimates of C demand for chaetognaths
For C5, the small deficit is within the margin of error would be offset by underestimates for euphausiids. In
for C demand. While flux distributions would be more addition, given our very conservative assumption of
complicated in an actual trophic network, for example no carnivorous feeding by euphausiids and many other
involving small and large microzooplankton and some use groups, we more likely underestimate total carnivore C
of detritus (Landry et al., 2020b), this simple analysis indi- demand. An advective subsidy would be another way to
cates that trophic flows from locally generated production meet the zooplankton production shortfall.
in the oceanic GoM in May are adequate to meet the C Kelly et al. (in review) argue for the importance of a
demands of suspension feeders, with most coming from lateral advective organic input from the coastal margins
predation on microzooplankton. to explain mass balance deficits of N export in the open-
From Table VI, the challenge for reconciling trophic ocean GoM. Mechanistically, material pulled northward
fluxes in the GoM appears to be the availability of by the loop current from the productive Campeche Bank
sufficient mesozooplankton production, here calculated of the Yucatan Peninsula (Merino, 1997; Melo Gonzalez
from Hirst and Sheader (1997), to satisfy carnivores. et al., 2000) could persist as a subsidy in the bodies of
Mesozooplankton production and carnivore C demand long-lived animals. Studies in the Northern GoM have
are balanced for C1, while C4 and C5 have production also documented evolving dynamics of phytoplankton

11
JOURNAL OF PLANKTON RESEARCH VOLUME 00 NUMBER 00 PAGES 1–15 2021

and zooplankton during transport in the Mississippi River


plume that leave depleted phytoplankton and a surplus of
zooplankton on the seaward end (Liu and Dagg, 2003;
Liu et al., 2005), where they can be transported into
the oceanic region by eddies. In the present study, back-
tracking of water parcels sampled during our Lagrangian
experiments shows a strong connectivity to the slope

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


margin region of the NE GoM several weeks prior to our
sampling for the C1 and C5 experiments with abundant
ABT larvae, but absent that connectivity for cycles C2–
C3 without larvae (Gerard et al., this issue). This suggest
that the NE slope margin may be both a hot spot of
adult spawning activity and a source region of lateral
organic enrichment to the larval ABT rearing habitat. It
may not be coincidental that the cycle (C5, Fig. 1) closest
to this area shows the greatest need for a lateral sub-
sidy of zooplankton to satisfy C demands of carnivores Fig. 4. Bulk d15 N and d13 C isotopic values for mesozooplankton size
while locally generated and balanced trophic fluxes are classes in the GoM. Uncertainties are standard errors of mean estimates.
sufficient to meet C demands of both suspension feeders
and carnivores for water parcel C1 in the central oceanic
GoM. Shropshire et al. (this issue) have further surmised
that a nutritional subsidy from advected zooplankton
may be important for larval ABT survival and point to
the shelf-slope margin as the optimal spawning areas
for minimizing starvation and predation risks of early
larvae.
Dorado et al. (2012) found no difference in δ 13 C val-
ues for zooplankton collected in neritic versus oceanic
waters in the northern GoM (both averaging −19.8 to
−19.9 ± 0.9 SD), despite very substantial δ 13 C differ-
ences in the POM from these two areas. Our similar val-
ues among size classes (Fig. 4) agree with their conclusion
of a common marine-based carbon source for the region.
In terms of trophic structure, the differences in percent
carnivorous feeders between large and small zooplankton Fig. 5. Day and nighttime estimates of percent carnivorous mesozoo-
plankton from net tows done during experimental cycles in the GoM.
are consistent with the inference of a 0.5–0.6 trophic Estimates for the small (<1-mm) size fraction are based on relative
step increase from δ 15 N isotopic enrichment (Figs. 4 and abundances. Estimates for the large (>1-mm) size fraction are based on
5). Despite no difference in δ 13 C, Dorado et al. (2012) relative DW.
reported significantly higher δ 15 N values for neritic versus
oceanic zooplankton (5.5 ± 1.1 and 2.8 ± 1.4, respec-
CONCLUSIONS
tively, for June and July collections), which they attributed
to the lower isotopic values of production from nitrogen Despite a habitat bordered closely by land masses on
fixers in the oceanic region. Our biomass-weighted mean almost all sides, mesozooplankton communities of the
values of δ 15 N from Fig. 4 are in the range of 4.5–4.7, so oceanic GoM exhibit generally similar characteristics to
closer to the previous δ 15 N values for coastal zooplankton. those of remote oligotrophic subtropical regions of the
In addition, nitrogen isotope mass balances for our exper- major oceans in terms of biomass, low grazing impact
imental cycles also revealed very little contribution from on phytoplankton and magnitude of active export from
diazotrophy during the times of our cruises (Knapp et al., respiration by diel migrators. Compared with averages
this issue). If these differences between our results and for HOT and BATS, however, zooplankton stocks are
Dorado et al. (2012) are indicative of a recurrent seasonal disproportionately high relative to contemporaneous pri-
pattern, they suggest a profound and rapid shift up in the mary production in the ABT larval rearing sites during
contribution of N2 fixation to biomass production as the the peak May season. Given the broader context of
ocean warms and further stratifies between May and July. process experiments that measured trophic fluxes and net

12
LANDRY AND SWALETHORP MESOZOOPLANKTON BIOMASS, GRAZING AND TROPHIC STRUCTURE

FUNDING
This study acknowledges BLOOFINZ-GoM Program support from
National Oceanic and Atmospheric Administration award
NA15OAR4320071 and U.S. National Science Foundation grants
OCE-1756517 and -1851558.

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


REFERENCES
Alemany, F., Quintanilla, L., Vélez-Belchí, P., García, A., Cortés, D.,
Rodríguez, J. M., DE Puelles, M. F., González-Pola, C. et al. (2010)
Characterization of the spawning habitat of Atlantic bluefin tuna
and related species in the Balearic Sea (western Mediterranean). Prog.
Oceanogr., 86, 21–38.
Fig. 6. Feeding requirements for mesozooplankton to meet carbon
demands for respiration and production in the GoM during Cycles 1–5 Al-Mutairi, H. and Landry, M. R. (2001) Active export of carbon and
experiments and the average day (D) and night (N) estimates for tows in nitrogen at station ALOHA by diel migrant zooplankton. Deep-Sea Res.
2017 and 2018. Yearly averages are divided into feeding requirements II., 48, 2083–2104.
to support suspension feeders (gray-scale shaded, below) and require- Avila-Alonso, D., Baetens, J. M., Cardenas, R. and De Baets, B. (2020)
ments to support carnivorous feeders (unshaded, top). Uncertainties are
standard errors of mean cruise estimates for the total community. Oceanic response to hurricane Irma (2017) in the exclusive economic
zone of Cuba and the eastern Gulf of Mexico. Ocean Dynm., 70,
603–619.
biomass changes in the ambient water column, growth- Babin, S. M., Carton, J. A., Dickey, T. D. and Wiggert, J.
grazing balances for phytoplankton were resolved with D. (2004) Satellite evidence of hurricane-induced phytoplankton
blooms in an oceanic desert. J. Geophys. Res., Oceans, 109. doi:
a slight positive, but statistically insignificant, net resid- 10.1029/2003JC001938.
ual, and trophic fluxes generated by local productivity
Bakun, A. and Broad, K. (2003) Environmental ’loopholes’ and fish
were found sufficient to satisfy C demand of suspension population dynamics: comparative pattern recognition with focus on
feeders. Microzooplankton were the major contributors to El Nino effects in the Pacific. Fish. Oceanogr., 12, 458–473.
these budgets. Other than the most remote experimental Bakun, A. (2006) Fronts and eddies as key structures in the habitat of
cycle, C1, which might be a model for how subtropical marine fish larvae: opportunity, adaptive response and competitive
ocean regions operate as trophically balanced systems, advantage. Sci. Mar., 70, 105–122.
experiments conducted closer to the GoM margin show Bakun, A. (2013) Ocean eddies, predator pits and bluefin tuna: implica-
a zooplankton production deficit, suggesting that lateral tions of an inferred “low risk-limited payoff” reproductive scheme pf
a (former) archetypical top predator. Fish Fish., 14, 424–438.
subsidy of zooplankton from the margins is important for
meeting carnivore C demand in the open-ocean region. Balch, W. M., Poulton, A., Drapeau, D. T., Bowler, B. C., Windecker, L.
A. and Booth, E. S. (2011) Zonal and meridional patterns of phyto-
plankton biomass and carbon fixation in the equatorial Pacific Ocean,
between 110◦ W and 140◦ W. Deep-Sea Res. II Top. Stud. Oceanogr., 58,
DATA ARCHIVING 400–416.
Data presented here have been submitted to the National Calbet, A., Atienza, D., Henriksen, C. I., Saiz, E. and Adey, T. R. (2009)
Zooplankton grazing in the Atlantic Ocean: a latitudinal study. Deep-
Oceanic and Atmospheric Administration’s (NOAA)
Sea Res II , 56, 954–963.
National Centers for Environmental Information (NCEI)
Cushing, D. H. (1990) Plankton production and yearclass strength in fish
data repository and will also be archived at BCO- populations: an update of the match/mismatch hypothesis. Adv. Mar.
DMO (Biological and Chemical Oceanography Data Biol., 26, 249–293.
Management Office) site https://www.bco-dmo.org/pro Dam, H. G., Zhang, X., Butler, M. and Roman, M. R. (1995) Meso-
gram/819631. zooplankton grazing and metabolism at the equator in the central
Pacific: implications for carbon and nitrogen fluxes. Deep-Sea Res II ,
42, 735–756.
ACKNOWLEDGEMENTS Décima, M., Landry, M. R. and Rykaczewski, R. (2011) Broad-scale
patterns in mesozooplankton biomass and grazing in the eastern
We gratefully acknowledge the capable leadership of John equatorial Pacific. Deep-Sea Res. II Top. Stud. Oceanogr., 58, 387–399.
Lamkin and Trika Gerard for the project overall, the cruise
Chief Scientists Estrella Malca (NF1704) and John Lamkin Décima, M., Landry, M. R., Stukel, M. R., Lopez-Lopez, L. and
(NF1802), and the captain and crew of the R/V Nancy Foster. Krause, J. W. (2016) Mesozooplankton biomass and grazing in the
We additionally thank Sarah Privoznic for her assistance in Costa Rica dome: amplifying variability through the plankton food
organizing the cruises, and Cameron Quackenbush, Juan web. J. Plankton Res., 38, 317–330.
Banuelos-Arriaga, Kelsey Fleming and Tabitha Hernandez for Domingues, R., Goni, G., Bringas, F., Muhling, B., Lindo-Atichati,
assistance with the field collections and laboratory analyses. D. and Walter, J. (2016) Variability of preferred environmental

13
JOURNAL OF PLANKTON RESEARCH VOLUME 00 NUMBER 00 PAGES 1–15 2021

conditions for Atlantic bluefin tuna (Thunnus thynnus) larvae in the Gulf production and grazing balances in the Costa Rica dome. J. Plankton
of Mexico during 1993-2011. Fish. Oceanogr., 25, 320–336. Res., 38, 366–379.
Dorado, S., Rooker, J. R., Wissel, B. and Quigg, A. (2012) Isotope Landry, M. R., Selph, K. E., Stukel, M. R., Swalethorp, R., Kelly, T.
baseline shifts in pelagic food webs of the Gulf of Mexico. Mar. Ecol. B., Beatty, J. L. and Quackenbush, C. R. (this issue) Microbial food
Prog. Ser., 464, 37–49. web dynamics in the oceanic Gulf of Mexico. J. Plankton Res..
Gerard, T., Lamkin, R. T., Kelly, T. B., Knapp, A. N., Laiz-Carrión, Landry, M. R., Selph, K. E., Taylor, A. G., Décima, M., Balch, W.
R., Malca, E., Selph, K. E., Shiroza, A. et al. (this issue) Bluefin M. and Bidigare, R. R. (2011) Phytoplankton growth, grazing and
larvae in Oligotrophic Ocean Foodwebs, investigations of nutrients to production balances in the HNLC equatorial Pacific. Deep-Sea Res. II ,

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


zooplankton: overview of the BLOOFINZ-Gulf of Mexico program. 58, 524–535.
J. Plankton Res.. Landry, M. R., Stukel, M. R. and Décima, M. R. (2020b) Food-
Hardy, A. C. (1924) The herring in relation to its animate environment, web fluxes support high rates of mesozooplankton respiration and
part I. the food and feeding habits of the herring with special production in the equatorial Pacific. Mar. Ecol. Prog. Ser.in press, 652,
reference to the east coast of England. Fish. Invest. Lond. Ser. 2, 7, 15–32. doi: 10.3354/meps13479.
1–53. Lindo-Atichati, D., Bringas, F., Goni, G., Muhling, B., Muller-Karger, F.
Hernández-León, S. and Ikeda, T. (2005) A global assessment of meso- E. and Habtes, S. (2012) Varying mesoscale structures influence larval
zooplankton respiration in the ocean. J. Plankton Res., 27, 153–158. fish distribution in the northern Gulf of Mexico. Mar. Ecol. Prog. Ser.,
Hirst, A. G. and Sheader, M. (1997) Are in situ weight-specific growth 463, 245–257.
rates body-size independent in marine planktonic copepods? A re- Liu, H. and Dagg, M. (2003) Interactions between nutrients, phyto-
analysis of the global synthesis and a new empirical model. Mar. Ecol. plankton growth, and micro- and mesozooplankton grazing in the
Prog. Ser., 154, 155–165. plume of the Mississippi River. Mar. Ecol. Prog. Ser., 258, 31–42.
Hjort, J. (1914) Fluctuations in the great fisheries of northern Europe Liu, H., Dagg, M. J., Wu, C.-J. and Chiang, K.-P. (2005) Mesozooplank-
viewed in the light of biological research. Rapp. P.-V. Réun. Cons. Int. ton consumption of microplankton in the Mississippi River plume,
Explor. Mer, 20, 1–228. with special emphasis on planktonic ciliates. Mar. Ecol. Prog. Ser., 286,
Ikeda, T. (1985) Metabolic rates of epipelagic marine zooplankton as a 133–144.
function of body size and temperature. Mar. Biol., 85, 1–11. Llopiz, J. K., Cowen, R. K., Hauff, M. J., Ji, R., Munday, P. L., Muhling,
Ikeda, T. (2014) Respiration and ammonia excretion by marine meta- B. A., Peck, M. A., Richardson, D. E. et al. (2014) Early life history and
zooplankton taxa: synthesis toward a global-bathymetric model. Mar. fisheries oceanography: new questions in a changing world. Oceanogr.,
Biol., 161, 2753–2766. 27, 26–41.
Isla, J. A., Llope, M. and Anadón, R. (2004) Size-fractionated meso- López, E. and Anadón, R. (2008) Copepod communities along an
zooplankton biomass, metabolism and grazing along a 50◦ N-30◦ S Atlantic Meridional transect: abundance, size structure and grazing
transect of the Atlantic Ocean. J. Plankton Res., 26, 1301–1313. rates. Deep-Sea Res. I , 55, 1375–1391.
Kelly, T. B., Knapp, A. N., Landry, M. R., Selph, K. E., Shropshire, Madin, L. P., Horgan, E. F. and Steinberg, D. K. (2001) Zooplankton at
T. A., Thomas, R. and Stukel, M. R. (in review) Lateral advection the Bermuda Atlantic time series study (BATS) station: diel, seasonal
supports nitrogen export in the oligotrophic open-ocean Gulf of and interannual variation in biomass, 1994-1998. Deep-Sea Res. II , 48,
Mexico. Nat. Commun.. 2063–2082.
Kleppel, G. S. and Pieper, R. E. (1984) Phytoplankton pigments in the Melo Gonzalez, N., Muller-Karger, F. E., Cerdeira Estrada, R. S., Perez,
gut contents of planktonic copepods from coastal waters off Southern R., Victoria, I., Cardennas Perez, P. and Arenal, I. M. (2000) Near-
California. Mar. Biol., 78, 193–198. surface phytoplankton distribution in the western intra-Americas sea:
the influence of El Nino and weather events. J. Geophys. Res., 105,
Knapp, A. N., Thomas, R., Stukel, M. R., Kelly, T. B., Landry, M. R., 14029–14043.
Selph, K. E., Malca, E., Gerard, T. et al. (this issue) Constraining the
sources of nitrogen fueling phytoplankton and food webs in the Gulf Minawaga, M. and Wada, E. (1984) Stepwise enrichment of 15 N along
of Mexico using nitrogen isotope budgets. J. Plankton Res.. food chains: further evidence and the relation between δ 15 N and
animal age. Geochim. Cosmochim. Acta, 48, 1135–1140.
Landry, M. R., Al-Mutairi, H., Selph, K. E., Christensen, S. and
Nunnery, S. (2001) Seasonal patterns of mesozooplankton abundance Merino, M. (1997) Upwelling on the Yucatan shelf: hydrographic
and biomass at station ALOHA. Deep-Sea Res. II , 48, 2037–2061. evidence. J. Mar. Syst., 13, 101–121.
Landry, M. R., Beckley, L. E. and Muhling, B. A. (2019) Climate Moriarty, R. and O’Brien, T. D. (2013) Distribution of mesozooplank-
sensitivities and uncertainties in food-web pathways supporting larval ton biomass in the global ocean. Earth Syst. Sci. Data, 5, 45–55.
bluefin tuna in subtropical oligotrophic oceans. ICES J. Mar. Sci., 76, Moriarty, R., Buitenhuis, E. T., Le Quére, C. and Gosselin, M.-P. (2013)
359–369. Distribution of known macrozooplankton abundance and biomass in
Landry, M. R., Hood, R. R. and Davies, C. H. (2020a) Mesozoo- the global ocean. Earth Syst. Sci. Data, 5, 241–257.
plankton biomass and temperature-enhanced grazing along a 110◦ E Muhling, B. A., Lamkin, J. T. and Roffer, J. T. (2010) Predicting
transect in the eastern Indian Ocean. Mar. Ecol. Prog. Ser., 649, 1–19. the occurrence of Atlantic bluefin tuna (Thunnus thynnus) larvae in
Landry, M. R., Ohman, M. D., Goericke, R., Stukel, M. R. and Tsyrke- the northern Gulf of Mexico: building a classification model from
vich, K. (2009) Lagrangian studies of phytoplankton growth and archival data. Fish. Oceanogr., 19, 526–539.
grazing relationships in a coastal upwelling ecosystem off Southern Muhling, B. A., Lee, S.–. K., Lamkin, J. T. and Liu, Y. (2011) Predict-
California. Prog. Oceanogr., 83, 208–216. ing the effects of climate change on bluefin tuna (Thunnus thynnus)
Landry, M. R., Selph, K. E., Décima, M., Gutiérrez-Rodríguez, A., spawning habitat in the Gulf of Mexico. ICES J. Mar. Sci., 68,
Stukel, M. R., Taylor, A. G. and Pasulka, A. L. (2016) Phytoplankton 1051–1062.

14
LANDRY AND SWALETHORP MESOZOOPLANKTON BIOMASS, GRAZING AND TROPHIC STRUCTURE

Owens, N. J. P. and Rees, A. P. (1989) Determination of nitrogen-15 at biogeochemical cycling. Global Biogeochem. Cycles, 26, GB1004. doi:
sub-microgram levels of nitrogen using automated continuous-flow 10.1029/2010GB004026.
isotope ratios mass spectrometry. Analyst, 114, 1655–1657. Straile, D. (1997) Gross growth efficiencies of protozoan and meta-
Roman, M. R., Adolf , H. A., Landry, M. R., Madin, L. P., Steinberg, zoan zooplankton and their dependence of food concentration,
D. K. and Zhang, X. (2002) Estimates of oceanic mesozooplankton: a predator-prey weight ratio, and taxonomic group. Limnol. Oceanogr.,
comparison using the Bermuda and Hawaii time-series data. Deep-Sea 42, 1375–1385.
Res. II , 49, 175–192. Strickland, J. D. H. and Parsons, T. R. (1972) A Practical Handbook of
Roman, M. R., Dam, H. G., Gauzens, A. L. and Napp, J. M. (1993) Seawater Analysis, Fisheries Research Board, Canada Ottawa.

Downloaded from https://academic.oup.com/plankt/advance-article/doi/10.1093/plankt/fbab008/6164925 by guest on 10 March 2021


Zooplankton biomass and grazing at the JGOFS Sargasso Sea time Sun, D. and Wang, C. (2017) Latitudinal distribution of zooplankton
series station. Deep-Sea Res. I , 40, 883–901. communities in the western Pacific along 160◦ E during summer 2014.
Selph, K. E., Swalethorp, R., Stukel, M. R., Kelly, T. B., Knapp, A. N., J. Mar. Syst., 169, 52–60.
Fleming, K., Hernandez, T. and Landry, M. R. (in press) Phytoplank- Taylor, A. G., Landry, M. R., Selph, K. E. and Yang, E. J. (2011)
ton community composition and biomass in the open-ocean Gulf of Biomass, size structure and depth distributions of the microbial
Mexico. J. Plankton Res.. community in the eastern equatorial Pacific. Deep-Sea Res. II , 58,
Sheridan, C. C. and Landry, M. R. (2004) A 9-year increasing trend 342–357.
in mesozooplankton biomass at the Hawaii Ocean time-series station Taylor, A. G., Landry, M. R., Freibott, A., Selph, K. E. and Gutiérrez-
ALOHA. ICES J. Mar. Sci., 61, 457–463. Rodríguez, A. (2016) Patterns of microbial community biomass, com-
Shiroza, A., Malca, E., Lamkin, J. T., Gerard, T., Landry, M. R., position and HPLC diagnostic pigments in the Costa Rica upwelling
Stukel, M. R., Laiz-Carrión, R. and Swalethorp, R. (this issue) Active dome. J. Plankton Res., 38, 183–198.
prey selection in developing larvae of Atlantic Bluefin tuna (Thunnus Valencia, B., Landry, M. R., Décima, M. and Hannides, C. C. S. (2016)
thynnus) in spawning grounds of the Gulf of Mexico. J. Plankton Res.. Environmental drivers of mesozooplankton biomass variability in
Shropshire, T., Morey, S. L., Chassignet, E., Karnauskas, M., Coles, the North Pacific Subtropical Gyre. Eur. J. Vasc. Endovasc. Surg., 121,
V. J., Malca, E., Laiz-Carrión, Fiksen, O. et al. (this volume) Trade- 3131–3143.
offs between risks of predation and starvation in larvae make the Valencia, B., Décima, M. and Landry, M. R. (2018) Environmental
shelf break an optimal spawning location for Atlantic Bluefin tuna. effects on mesozooplankton size structure and export flux at station
J. Plankton Res.. ALOHA, North Pacific subtropical gyre. Global Biogeochem. Cycles, 32,
Steinberg, D. K., Carlson, C. A., Bates, N. R., Goldthwait, S. A., Madin, 289–305.
L. P. and Michaels, A. F. (2000) Zooplankton vertical migration and Yingling, N., Kelly, T. B., Selph, K. E., Landry, M. R., Knapp, A. N.,
the active transport of dissolved organic and inorganic carbon in the Kranz, S. A. and Stukel, M. R. (this issue) Taxon-specific phytoplank-
Sargasso Sea. Deep-Sea Res. I , 47, 137–158. ton growth, nutrient limitation, and light limitation in the oligotrophic
Steinberg, D. K. and Landry, M. R. (2017) Zooplankton and the ocean Gulf of Mexico. J. Plankton Res..
carbon cycle. Ann. Rev. Mar. Sci., 9, 413–444. Zhang, X., Dam, H. G., White, J. R. and Roman, M. R. (1995)
Steinberg, D. K., Lomas, M. W. and Cope, J. S. (2012) Long- Latitudinal variations in mesozooplankton grazing and metabolism
term increase in mesozooplankton biomass in the Sargasso Sea: in the central tropical Pacific during the US JGOFS EqPac study.
linkage to climate and implications for food web dynamics and Deep-Sea Res. II Top. Stud. Oceanogr., 42, 695–714.

15

You might also like